Synthesis and Surface Functionalization of Hydride ... - ACS Publications

Apr 10, 2017 - silicon, bulk Ge possesses a smaller band gap (0.67 vs 1.1 eV at. 300 K),6 larger ... electron and hole mobility (≤3900 vs ≤1500 cm...
0 downloads 0 Views 4MB Size
Subscriber access provided by HACETTEPE UNIVERSITESI KUTUPHANESI

Article

Synthesis and Surface Functionalization of Hydride-terminated Ge Nanocrystals Obtained from the Thermal Treatment of ‘Ge(OH)2’ Morteza Javadi, Darren Picard, Regina Sinelnikov, Mary Alvean Narreto, Frank Anthony Hegmann, and Jonathan G. C. Veinot Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b00358 • Publication Date (Web): 10 Apr 2017 Downloaded from http://pubs.acs.org on April 16, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Synthesis and Surface Functionalization of Hydrideterminated Ge Nanocrystals Obtained from the Thermal Treatment of ‘Ge(OH)2’

Morteza Javadi,1 Darren Picard,1 Regina Sinelnikov,1 Mary Alvean Narreto,2 Frank A. Hegmann,2 and Jonathan G. C. Veinot1*

The synthesis of germanium nanocrystals (GeNCs) with well-defined surface chemistry is of considerable interest because of their potential applications in the optoelectronic, battery, and semiconductor industries. Modifying and tailoring GeNC surface chemistry provides an avenue by which reactivity, environmental compatibility (e.g., solubility, resistance to oxidation), and electronic properties may be tailored. Hydride-terminated GeNCs (H-GeNCs) are of particular interest because the reactivity of surface Ge-H bonds toward alkenes and alkynes via hydrogermylation affords the potential for convenient modification, however these reactions and their scope have not been widely explored. This report describes a straightforward route for preparing a GeNC/GeO2 composite via disproportionation of heretofore-unexplored Ge (II) oxide-based precursor from which the H-GeNCs were freed by subsequently chemical etching. The H-GeNCs were derivatized using a series of hydrogermylation approaches (i.e., thermallyactivated, radical-initiated, and borane-catalyzed). The presented findings indicate surface functionalization occurs under all conditions investigated; however the nature of surface species (i.e., monolayers vs. multilayers) and surface coverage varies depending upon the conditions employed.

1

Department of Chemistry, University of Alberta, 11227 Saskatchewan Drive, Edmonton, Alberta, Canada. E-mail: [email protected]; Fax: +1-780-492-8231; Tel: +1-780-492-7206 2 Department of Physics, University of Alberta, Edmonton, Alberta T6G 2E1, Canada.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Introduction General interest in germanium came about with the realization of semiconductor-based electronics in the second half of 20th century.1 Transistors, which are essential components in the electronic devices upon which modern society relies, were first made using germanium.2,3 While Ge lost out to ultra-pure silicon, it remains an important material for optoelectronic applications.1 In the early 1990s the surprising discovery of visible photoluminescence (PL) from silicon and germanium nanostructures opened new possibilities.4,5 Despite being in Group 14 germanium differs substantially from its more frequently studied periodic congener silicon. Compared to silicon, bulk Ge possesses a smaller band gap (0.67 vs. 1.1 eV at 300 K),6 larger Bohr-exciton radius (24.3 vs. 4.9 nm),5,7 higher electron and hole mobility (≤3900 vs. ≤1500 cm2/V·s),8 as well as greater capacity for and diffusivity of ions (e.g., Li+).9 In the context of these characteristics, comparatively large GeNCs and related structures (e.g., oxide-embedded GeNCs) are expected possess favorable chemical, optical and electronic properties useful in applications such as solar cells,10 biological imaging,11,12 Bragg reflectors,10 light-emitting diodes, non-volatile memory devices,13–17 as well as battery electrode materials.9,18 A variety of procedures for preparing GeNCs have appeared; some highlights include: laser ablation19 and ball milling of bulk Ge,20 plasma pyrolysis of GeH4,21 solution-phase reduction of Ge(II) and Ge(IV) precursors,22–24 metathesis of Ge-based Zintl salts,25–28 and thermal decomposition of organogermane precursors (conventional,29–31 and microwave heating32). While other germanium halide solution-phase reduction protocols have been reported,33 a ground-breaking report by Klimov et al. showed the first example of colloidal GeNCs derived from solution reduction that photoluminesce in the infrared spectral region via a band-gap transition.10 The authors also effectively showed size dependence of GeNC optical properties that arise from quantum confinement.10 Despite these important advances, developing methods that afford GeNCs with well-defined surface chemistry remains an important target if their full potential is to be realized.33,34 Solid-state syntheses of oxide-embedded SiNCs via thermolysis of substoichiometric oxide-based polymeric precursors have proved very effective.35,36 These procedures provide control of NC dimension and shape, and afford NCs whose surface chemistry can be tailored subsequent to liberation.35,37,38 A similar approach using sol-gel derived organic functionalized

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

germanium rich precursors (i.e., [(RGeO1.5)n]) showed promise.39 Specifically, a phenyl substituted sol-gel polymer [(PhGeO1.5)n] was thermally treated to yield GeNCs embedded in GeO2. This early investigation showed Ge was formed via two pathways (i.e., disproportionation of substoichiometric oxides and direct reduction of oxide by hydrogen carrier gas). The GeNCs were freed from the oxide matrix upon etching with warm water, however size distributions were broad.39 Subsequent investigations aimed at exploring the role of the organic substituents indicated carbon contamination restricted the liberation of GeNCs from oxide matrix limiting the utility of this approach.40 In the present contribution we describe a facile method for preparing GeNCs from a carbon-free germanium-rich oxide precursor derived from commercial GeO2 powder. Thermal processing of this precursor in inert atmosphere yields GeO2-embedded GeNCs that are readily liberated with hydride surfaces (i.e., Ge-H) via HF etching and subsequently functionalized via hydrogermylation.

Experimental Reagents and Materials. Germanium dioxide powder (GeO2, 99.9%) was purchased from Eagle-Picher. Hypophosphorous acid solution (50 wt. % in H2O), sodium hydroxide pellets, 1dodecene (97%), Chloroform-d (CDCl3, 99.8 atom % D), 1-dodecene (97%), 2,2'-azobis(2methylpropionitrile) (AIBN, 98%), benzoyl peroxide (98%), and borane tetrahydrofuran (BH3⋅THF 1M) as well as reagent grade methanol, toluene, and ethanol were purchased from Sigma-Aldrich. Electronics grade hydrofluoric acid (HF, 49% aqueous solution) was purchased from J. T. Baker. Hydrochloric acid (36.5 - 38%) and ammonium hydroxide (28 - 30%) were purchased from Caledon Lab. Ultrapure H2O (18.2 MΩ/ cm) purified in a Barnstead Nanopure Diamond purification system was used in all reactions. Molecular sieves (type 4 Å) were purchased from Sigma-Aldrich and activated in vacuum oven prior to use. Unless otherwise indicated reagents were used as received. Synthesis of germanium (II) hydroxide, ‘Ge(OH)2’. A modified literature procedure was used to prepare ‘Ge(OH)2’.41–46 In a typical preparation, 1.0 gram of germanium dioxide powder was dissolved in 7.0 mL of a freshly prepared aqueous NaOH (~17 M). Upon slow addition of 24 mL

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

aqueous HCl (6 M), GeO2 powder precipitated and then dissolved to give a clear acidic solution. Subsequently, 7.5 mL of aqueous 50 % H3PO2 was added all at once and the resulting colourless solution was refluxed under an argon atmosphere. After 5.5 hours the condenser was removed and concentrated aqueous NH4OH was added dropwise to yield a brown precipitate of ‘Ge(OH)2’ (Caution: This reaction is extremely exothermic). The brown precipitate was recovered

by vacuum

filtration

and

washed

three

times

with

ca.

100

mL of

deionized/deoxygenated water. Finally the solid was dried vacuum oven at 50 °C for 12 hours and stored in a vial until needed. Typical yields exceed 85%. Thermal processing of germanium (II) hydroxide. Thermal processing was performed using a Lindberg/Blue tube furnace in flowing argon (15 mL/min). Typically, 0.5 g of ‘Ge(OH)2’ powder was transferred to a quartz reaction boat, placed in the furnace and heated to 400 °C at 20 °C/min where it remained for 60 minutes. GeNCs in a germanium dioxide matrix (GeNCs/GeO2) was obtained as a black/brown powder (ca. 0.5 g), ground using an agate mortar and pestle, and stored under ambient atmosphere for further use and characterization. GeNC Isolation. Hydride-terminated GeNCs (H-GeNCs) were liberated from the oxide matrix upon etching of GeNCs/GeO2 composite with alcoholic HF. The fine powder obtained from grinding the product of thermal processing (~50 mg) was magnetically stirred in 1.0 mL of 1:1 solution of ethanol and water for 15 min in a polyethylene terephthalate (PET) beaker. Subsequently, 0.5 mL of HF acid (49% aqueous solution) was added dropwise. (Caution! HF must be handled with extreme care and in accordance with local regulations.) After stirring the mixture for 15 min, three portions of ca. 5 mL toluene were added to extract the H-GeNCs. The cloudy toluene layer was centrifuged at 3000 rpm to yield a black precipitate. This precipitate was dispersed in toluene containing activated 4Å molecular sieves (ca. 1 g) and gently agitated for approximately one minute to remove trace moisture. The suspension was transferred to centrifuge tubes and the black precipitate (H-GeNCs) was recovered upon centrifuging at 3000 rpm and immediately functionalized using procedures outlined below. Thermally induced hydrogermylation of H-GeNC. The black precipitate (i.e., H-GeNCs.) isolated from the etching of the GeNCs/GeO2 composite (ca. 50 mg) was dispersed in 5.0 mL 1dodecene, transferred to a Schlenk flask attached to an argon charged double manifold and three freeze-pump-thaw cycles were performed. Subsequently, the cloudy deoxygenated reaction

ACS Paragon Plus Environment

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

mixture was heated to 190 °C with stirring. After approximately 15 h the reaction mixture becomes transparent and takes on an orange-brown appearance. The resulting dodecane functionalized GeNCs (dodecyl-GeNCs) were isolated (ca. 15 mg) and purified as outlined below. Luminescent dodecyl-GeNCs were prepared by introducing small quantity of I2 (ca. 3 mg) to reaction flask prior to heating. Radical initiated hydrogermylation of H-GeNC. H-GeNCs (from etching ca. 50 mg of GeNCs/GeO2 composite) were dispersed in 5 ml of a of 1-dodecene and the radical initiator of choice (i.e., 10 mg AIBN or 15 mg BP) was added. The mixture was transferred to a Schlenk flask attached to an argon charged double manifold and exposed to three freeze-pump-thaw cycles. Subsequently, the argon filled reaction flask was heated (60 °C for AIBN or 85 °C for BP) and stirred for 15 hours to yield an orange-brown transparent suspension of crude dodecylGeNCs (ca. 15 mg) that were subjected to purfication procedures outlined below. Borane catalyzed surface hydrogermylation. H-GeNCs (from etching ca. 50 mg of GeNCs/GeO2 composite) were dispersed in 5 ml of 1-dodecene. The reaction mixture was transferred to a Schlenk flask attached to an argon charged double manifold and exposed to three freeze-pump-thaw cycles. Subsequently, 560 µL of BF3THF (1M) (i.e., 2.5 mol % of 1dodecene) was added via syringe and the argon filled reaction flask was stirred at room temperature (23 °C) for 15 hours to yield transparent orange-brown suspensions of crude dodecyl-GeNCs (ca. 15 mg) that were subjected to purfication procedures outlined below. Isolation and purification of 1-dodecyl-GeNCs. The same solvent/antisolvent precipitation procedure was used to purify dodecyl-GeNCs obtained from the functionalization procedures noted above. The transparent orange-brown suspensions of crude dodecyl-GeNCs obtained from the functionalization procedures were transferred to a polypropylene centrifuge tubes and toluene (5 mL) with ca. 45 mL of a 1:1 ethanol and methanol (v:v) antisolvent were added. The resulting cloudy brown suspension was centrifuged at 12000 rpm for 30 minutes to yield brown precipitate and transparent colorless supernatant. The supernatant was discarded. The brown precipitate was resuspended with sonication in a minimum volume of toluene (ca. 2 ml). Subsequently, methanol (ca. 50 mL) was added to yield a cloudy suspension and the mixture was centrifuged at 12000 rpm for 30 min to yield a brown solid and clear colourless supernatant. The supernatant was decanted and discarded. The suspension/centrifugation procedure was repeated twice after which

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the precipitate consisting of purified dodecyl-GeNCs was dispersed in ca. 5 mL of dry toluene and stored in a vial for material characterization. Homo-oligomerization of 1-dodecene. To evaluate if 1-dodecene oligomers form during the functionalization procedures investigated in the absence of GeNCs, samples were prepared by treating 5 mL of neat 1-dodecene using the identical conditions used for the functionalization protocols of choice (i.e., no GeNCs). Material Characterization and Instrumentation. Fourier-Transform Infrared Spectroscopy (FT-IR) was performed using a Nicolet Magna 750 IR spectrophotometer. Samples were drop cast from a toluene suspension containing material in question. X-ray Powder Diffraction (XRD) was performed using an INEL XRG 3000 X-ray diffractometer equipped with a Cu-Kα radiation source (λ = 1.54 Å) and CPS-120 detector. Bright field transmission electron microscopy (TEM) and energy dispersive X-ray (EDX) analyses were performed using a JEOL-2010 (LaB6 filament) electron microscope with an accelerating voltage of 200 kV. High resolution TEM (HRTEM) imaging was performed on JEOL-2200FS TEM instrument with an accelerating voltage of 200 kV. TEM samples of GeNCs were prepared by drop-coating of 1-3 drops of a dilute toluene solution containing GeNCs of choice onto a holey carbon coated copper grid (300 mesh, Electron Microscopy Science) and the solvent was removed in vacuo. Bright field TEM and HRTEM images were processed using ImageJ software (version 1.48 v). Particle size distributions were determined using measurements of 200-300 NCs. Raman spectroscopy was performed using a Renishaw inVia Raman microscope equipped with a 514 nm diode laser and a power of 3.98 mW on the sample. Samples were prepared by mounting the suspension on gold coated glass. X-ray photoelectron spectra were acquired in energy spectrum mode at 210 W, using a Kratos Axis Ultra X-ray photoelectron spectrometer. X-Ray source was Al (Mono) Kα line. Energy is 1486.6 eV. The probing area is about 1 x 2 mm2. Samples were prepared as films drop-cast from toluene solution/suspension onto a copper foil substrate. Binding energies were calibrated using the C 1s peak as a reference (284.8 eV). CasaXPS Version 2.3.5 software was used. Peak fitting was performed after background subtraction (Shirley-type). The highresolution Ge 3d region of the XP spectra has been collected for all samples investigated and were fit to Ge 3d3/2/Ge 3d5/2 partner lines, with spin-orbit splitting fixed at 0.6 eV, and the Ge 3d3/2/Ge 3d5/2 intensity ratio was set to 0.67. Nanostructured-assisted laser desorption/ionization mass spectroscopy (NALDI-MS) was performed in positive/negative reflection mode using a

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Bruker Daltonics UltrafleXtreme MALDI TOF/TOF mass spectrometer. Samples were prepared by spotting ~1 µL of GeNC toluene suspension in question onto a Bruker Daltonics NALDI target and air-dried. Nuclear magnetic resonance spectroscopy (NMR) spectra were obtained using a Varian Unity INova Console 500 MHz NMR spectrometer. A concentrated solution (ca. 3 mg/mL) of dodecyl-GeNCs in CDCl3 was used to collect the 1HNMR spectra. FID files were processed using Nuts NMR data processing software. Photoluminescence (PL) measurement and Instrumentation. An ultrafast laser at 800 nm excitation source with 55 fs pulse width and 1 kHz repetition rate was employed. A fluence of ~340 µJ was used to excite solutions of GeNCs and the PL was collected at the back of the cuvette using lenses and optical fiber onto the entrance slit of the monochromator (Triax 180, Horiba), which has a grating at the NIR region with a blaze at 1500 nm. The output at the exit slit is refocused to a thermoelectronically-cooled InGaAs amplified photodetector (PDA10DT, Thorlabs) which has a wavelength range of 0.9  2.57 m set at 1MHz bandwidth with 70 dB gain. The signal is then processed through a DSP lock-in amplifier (SR830, Stanford Research). The PL spectra were plotted by a Labview program which also controls the monochromator.

Results and Discussion Germanium (II) hydroxide, commonly referred to hydrous germanium (II) oxide, has an uncertain stoichiometry.42 The ‘Ge(OH)2’ precursor employed herein to prepare GeNCs was precipitated from a hot aqueous Ge(II) solution (Scheme 1).41,44,47 Scheme 1. Preparation of dodecane functionalized GeNCs. (1), Thermal processing of ‘Ge(OH)2’ at T = 400 °C in Ar, (2), Liberation of H-GeNCs via HF etching, (3), Functionalization/surface modification of GeNCs through thermally-activated, radical-initiated, or borane-catalyzed hydrogermylation.

Figure 1a shows XRD patterns obtained for ‘Ge(OH)2’ and composites prepared upon thermal processing at 400 °C (for 1 h under Ar). The diffraction pattern of the ‘Ge(OH)2’

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

precursor shows broad features at ca. 2θ = 25° - 35° and 45° - 50° consistent with the presence of amorphous Ge.48 In addition, low intensity, sharp reflections at 2θ = 28.7, 37.4, 56.7, 59.3, and 72.3° appear that are assigned to rutile GeO2 (P42/mnm (136)).49 Heating the ‘Ge(OH)2’ at 400 °C under Ar provides a dark-brown/black powder the XRD pattern of which (Figure 1a) shows broadened reflections characteristic of diamond cubic crystalline Ge at 27.3, 45.3, 53.7, 66.1, 72.9 and 83.8° consistent with nanocrystalline domains of germanium being present. DebyeScherrer analysis based on the {111} crystallographic planes of the present GeNCs (assuming the dimensionless shape factor is 0.94) indicates crystal sizes of 6.1 nm upon the thermal processing of brown precursor at 400 °C.

Figure 1 Characterization of ‘Ge(OH)2’ before and after thermal processing at 400oC (for 1 h under Ar) (a) X-ray powder diffraction patterns. Standard reflections of crystalline Ge are provided.

49,50

Rutile GeO2 reflections are

indicated using a *. (b) XP spectra of the Ge 3d spectral region. (c) Raman spectra.

The survey XP spectra of brown ‘Ge(OH)2’ before and after thermal processing at 400 ºC under Ar (Figure S2), show Ge, C and O as expected along with peaks in 300-600 eV related to Ge LMM.51,52 High resolution X-ray photoelectron spectroscopy data (Figure 1b) are consistent with a disproportionation process occurring during thermal processing that sees the Ge in ‘Ge(OH)2’ converted to Ge(0) and Ge(IV) in the GeNC/GeO2. The spectrum of unprocessed ‘Ge(OH)2’ ((Figure 1b bottom trace) shows a broad emission in the Ge 3d region that is dominated by a Ge(II) emission. After thermal processing (Figure 1b, top trace), two distinct emissions appear at 29 and 33 eV consistent with formation of Ge(0) and Ge(IV), respectively.53

ACS Paragon Plus Environment

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Raman spectroscopy allows detection/identification of Ge-Ge bonds. In addition, the shape and position of the Ge-Ge feature provides insight into the size and crystallinity of GeNCs.39,54,55 Asymmetric peaks at high frequencies featuring shoulders at low frequencies are routinely attributed more crystalline nanoparticles.56 Raman analyses of ‘Ge(OH)2’ and GeNC/GeO2 (Figure 1c) are consistent with amorphous Ge nuclei within the precursor transforming to nanocrystalline Ge domains with thermal processing. The spectrum of ‘Ge(OH)2’ shows a symmetric optical phonon at ca. 280 cm-1 that we attribute to small, noncrystalline Ge nuclei while the same analysis of GeNC/GeO2 shows an asymmetric feature at high frequencies (ca. > 290 cm-1) with a shoulder consistent with Ge nanodomain crystallization. FT-IR analyses provides some limited insight into the bonding within ‘Ge(OH)2’ and GeNC/GeO2 (Figure S1). The spectrum of ‘Ge(OH)2’ shows broad O-H asymmetric/symmetric stretching bands in the range of 2900 to 3700 cm-1, as well as a bending band at 1655 cm-1. In addition, a feature at 755 cm-1 is attributed to Ge-O stretching. When comparing the spectra of GeNC/GeO2 to that of ‘Ge(OH)2’ (Figure S1) we note the broad feature at 3400 cm-1 is dramatically diminished and the broad Ge-O absorbance at 755 cm-1 moves to higher energy (i.e., 860 cm-1) supporting the proposal that a Ge-O-Ge network formed. While there are numerous potential applications for GeNC/GeO2 composites,13–17 the focus of this contribution lies in the preparation and tailoring surface chemistry of freestanding colloidal GeNCs. Selective removal of the GeO2 matrix from the present GeNC/GeO2 composite may be achieved using a variety of chemical etching protocols (e.g., warm water (> 60 °C),39 basic solutions57). For the present study we chose to employ HF etching because this procedure is expected to yield H-GeNCs that possess surfaces amenable to hydrogermylation.35 The FT-IR spectrum of the liberated GeNCs (Figure 2a) shows two sharp features at ca. 2000 and 750 cm-1 consistent with Ge-H stretching and bending. Of important note, there is negligible indication of surface oxide at the sensitivity of the FT-IR method. Additional characterization of the H-GeNCs was precluded by their limited solvent compatibility and air sensitivity. Three complementary hydrogermylation approaches (i.e., thermally-activated, radicalinitiated, and borane-catalyzed) were explored to achieve alkyl surface functionalization of HGeNCs using 1-dodecene as a model substrate. For convenience the following discussion of alkyl-terminated GeNCs will focus upon those prepared using the thermally initiated reactions.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

FT-IR spectra of the resulting dodecyl-GeNCs are shown in Figure 2a (thermally-activated) and Figure S3 (radical-initiated, borane-catalyzed). In all cases we note the intensity of the Ge-H diminishes dramatically and intense absorptions at ca. 2850 - 2650 and 1460 cm-1 appear; we confidently attribute these new features to C–H stretching and bending modes of surface bonded aliphatic chain.58 Further evidence of alkyl moiety covalent surface attachment is provided by a Ge-C stretching absorption at 700 cm-1.59 Features at ca. 860 cm-1 are attributed to Ge-O stretching modes arising from oxidation that presumably occurs during workup following functionalization. Figures 2b and S4a show representative bright field transmission electron micrographs of dodecyl-GeNCs. The particles appear pseudospherical with an average diameter of 7.2 ± 0.9 nm. This is in agreement with Debye-Scherrer size estimation. TEM analyses of dodecylfunctionalized GeNCs prepared using radical-initiated, and borane-catalyzed hydrogermylation approaches show similar morphology and size distributions (Figure S4). HRTEM imaging (Figure 2b inset, and Figure S4) of dodecyl-GeNCs shows crystalline nano-domains with fringes separated by 0.33 nm that correspond to Ge (111) lattice spacing.60 Energy-dispersive X-ray analysis (EDX) (Figure 2c) confirms the presence of germanium, carbon and a small amount of oxygen. Raman spectra (Figure 2d) show an asymmetric feature at 290 cm-1 indicating the crystallinity of the germanium nanodomains is maintained throughout the etching of GeNC/GeO2 composites and functionalization. Moreover, dodecyl-functionalized GeNCs were characterized by XPS and XRD (Figure S5). Figure S5a shows a representative survey XP spectrum of dodecyl-GeNCs; there are features corresponding to the binding energies of Ge 3d, Ge 3p, Ge 3s, C 1s, and O 1s. Deconvolution of Ge 3d spectral region (Figure S5b), shows dodecyl-GeNCs are dominated by elemental Ge along with some Ge(IV) species consistent with partial oxidation (as shown in FT-IR). The XRD pattern of dodecyl-GeNCs in Figure S5c shows the clear broad reflection from nanocrystalline domains of germanium. The XRD pattern of dodecyl-GeNCs also confirms effective removal of GeO2 matrix.

ACS Paragon Plus Environment

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 2 Characterization of GeNCs obtained from the thermal treatment of ‘Ge(OH)2’ at 400 °C for 1 h in a flowing Ar atmosphere. (a) FT-IR spectra of hydride-terminated (black trace) and dodecyl-GeNCs obtained from thermal hydrogermylation (red trace). (b) Bright-field TEM micrograph of dodecyl-GeNCs. (c) and (d) EDX and Raman spectra of dodecyl-GeNCs obtained from thermal hydrogemylation.

The presented analyses are consistent with surface hydrogermylation proceeding under all conditions explored here; however, investigation of analogous surface hydrosilylation reactions on silicon nanocrystals indicates the nature of surface bonded species and surface coverage

are

method

dependent.38

In

this

context,

nanostructured-assisted

laser

desorption/ionization mass spectroscopy (NALDI-MS) was employed to interrogate the monomeric or oligomeric nature of surface bonded species. To isolate the role of GeNCs NALDI analysis was also performed on product mixtures obtained from the identical reaction conditions used for surface modification - in the absence of GeNCs no high molecular weight fragments were detected (See Figure S6). In all cases involving GeNCs no fragments corresponding to the molecular weight of 1dodecene (MW = 168.3) were detected at the sensitivity of the MS technique consistent with effective purification. The fragmentation patterns obtained for all dodecyl-GeNCs are complex regardless of the functionalization activation method employed (Figures 3 a-d). Despite this complexity, insight into the nature of the GeNC surface species can be obtained by considering peak series with mass labels of the same color; these series are separated by the mass of the dodecyl repeat unit (i.e., 168.3 m/z). Similar patterns have been attributed to ligand

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

oligomerization on NC surfaces.24 Samples prepared using thermally activated hydrogermylation (Figure 3a) show fragments containing multiple dodecyl repeat units consistent with there being a higher amount of oligomerization occurring on the GeNC surfaces. In addition, we note high m/z fragments with the characteristic Ge isotopic signature. Because the Ge-Ge linkage is the weakest of the surface bonds (i.e., Ge-Ge, 190-210; Ge-C, 255; and C-C, 292-360 KJ·mol-1),61 it is reasonable it can cleave preferentially to produce high m/z fragments including differing numbers of germanium atoms (Figure S7). Similar clusters of Ge atoms including various numbers of Ge atoms (such as Ge9 clusters) have been reported previously.62,63 GeNCs modified using AIBN and BP initiated hydrogermylation (Figure 3 b and c) also show fragmentation patterns associated with multiple dodecyl repeat units, and borane catalyzed hydrogermylation (Figure 3d) resulted in the comparatively low m/z (i.e., not exceeding 500 m/z) suggesting limited (even negligible) oligomerization.

ACS Paragon Plus Environment

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 3. Nanoassisted laser desorption ionization mass spectra of dodecyl-GeNCs functionalized using a) thermally-activated (inset: higher m/z showing dodecene repeat units) b) AIBN and c) BP radical-initiated, d) borane-catalyzed hydrogermylation. Inset spectra are higher magnification from larger m/z.

Proton nuclear magnetic resonance (1H NMR) spectroscopy provides an alternative method for interrogating surface bonded moieties. Figure 4 shows 1H NMR spectrum of dodecylGeNCs functionalized via thermally-activated hydrogermylation (See Figure S8 for 1H NMR spectra of dodecyl-GeNCs obtained using other methods). All dodecyl-GeNCs show the expected broad resonances arising from terminal methyl protons at ca. 0.9 ppm and a broad resonance methylene chain protons in the range of ca. 1.1−1.6 ppm. Broad structureless aliphatic CH signals indicative of multiple, surface bonded environments.58,64,65 In addition, no features

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

associated with alkene protons are detected at the sensitivity of the 1H NMR method consistent with 1-dodecyl moieties being tethered to the GeNC surfaces.66 For the present dodecyl-GeNCs, the integration ratios of the methylene (at 1.29) and methyl (at 0.9 ppm) proton signals were determined to be 6.72, 6.09, 6.53 and 6.86 for thermally induced, AIBN initiated, BP initiated and borane catalyzed hydrogermylation, respectively. These values are consistent with the expected surface dodecyl moiety or dodecyl oligomers.65,67

Figure 4. 1H NMR spectrum of dodecyl-GeNCs obtained from thermally activated hydrogermylation in CDCl3 containing 0.01% (v/v) TMS. Integration of dodecyl methyl protons (3H) to TMS protons (12H) are shown. Methyl and chain methylene protons denoted by a and b, respectively. Residual solvent impurities (H2O or HOD at 1.5 ppm and toluene CH3 at 2.34 ppm)68,69 are denoted by (∗).

1

H NMR also offers a method to estimate NC surface coverage. Evaluating a ratio of the

integrated peak areas of the surface organic groups and that of an internal standard (i.e., TMS 0.01%) provides an approximation for amount of surface ligand.65 The results from these calculations (Table S1 in SI) indicate that, among different hydrogermylation protocols applied here, thermally induced functionalization provide highest degree of surface coverage (207%; NB: Based upon 1-dodecene monolayer coverage.). AIBN and BP radical initiate reactions, as well as

those catalyzed by borane provide lower surface coverage (i.e., 62%, 84% and 56%, respectively).58,65 These data are consistent with the NALDI-MS analyses noted above that indicated the likelihood of surface oligomerization decreased in the order of thermally activated,

ACS Paragon Plus Environment

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

radical initiation, and borane catalyzed hydrogermylation (i.e., a higher degree of surface oligomerization will lead to artificially high surface coverage). Having prepared dodecyl-GeNCs from the thermal processing of ‘Ge(OH2)’, we endeavoured to investigate their optical response (i.e., PL). As a result of the small band gap of Ge, electron-hole recombination in GeNCs is expected to exhibit PL in the NIR or IR spectral regions. However, reports are often contradictory regarding the source of PL, critical NC size, expected PL wavelength, quantum yield, effect of oxides and other impurities.10,33,39,40,70 Adding further complexity, germanium oxides can often show visible PL.7,50 There are also numerous examples showing UV-Vis emitting GeNCs,7,71,72 however it has been suggested that these emitters likely show PL resulting from surface species (e.g., oxides) rather than quantum confinement of GeNCs.34,73–75 There are surprisingly few examples of colloidal GeNCs showing NIR or IR PL.10,70,73 Despite the average size of the functionalized GeNCs investigated here being smaller than the Bohr-exciton radius of Ge (i.e., 24.3),5,7 they do not show detectable photoluminescence (PL). We hypothesize this may result from dark surface defects (e.g., impurities, vacancies, surface oxides, dangling bonds and etc.)76 arising from the HF liberation protocol that remain after hydrogermylation, that provide non-radiative pathways.70,77,78 In an effort to passivate potential (as of yet unidentified) defects on the GeNC surfaces we introduced a trace quantity of I2 into the hydrogermylation reaction mixture. GeNCs prepared in this way show PL at 986 nm (See Figure S9a). Survey XP spectra of the resulting dodecyl-GeNCs (Figure S9b) show evidence of only Ge, C and O: no iodine was detected at the sensitivity of the XPS method (ca. 1 – 3 atomic %, I 3d at 619 eV Figure S9b inset). At this time, the role of I2 in inducing PL remains unclear and is the subject of ongoing investigation.

Conclusion ‘Ge(OH2)’ was synthesized using a commercial GeO2 powder and employed as a precursor for H-GeNCs. This procedure exploits a disproportionation reaction to yield GeNCs embedded in a GeO2 matrix that are readily liberated with hydride surface moieties by HF etching. The resulting H-GeNCs provide a reactive platform that can be further modified using

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

thermal, radical initiated and borane-catalyzed hydrogermylation. NALDI-MS and 1H NMR analysis indicated oligomerization and surface coverage occurring by thermally induced method compared to other hydrogermylation procedures. Dodecane functionalized GeNCs prepared by proper defect removing approach (i.e. introducing I2), possess optical response in IR region.

Acknowledgments The authors recognize generous continued funding from the Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant Program and the “Alberta/Technische Universität München Graduate School for Functional Hybrid Materials (ATUMS)” supported by the NSERC CREATE and DFG IRTG Programs. The University of Alberta Faculty of Graduate Studies, Faculty of Science, and Department of Chemistry also provided financial support. G. Popowich and W. Moffat are thanked for assistance. Dr. Y. Khaniani and the members of the Veinot research team are also thanked for useful discussions.

References (1)

Haller, E. E. Germanium: From Its Discovery to SiGe Devices. Mater. Sci. Semicond. Process. 2006, 9 (4–5), 408–422.

(2)

Patton, G. L.; Iyer, S. S.; Delage, S. L.; Tiwari, S.; Stork, J. M. C. Silicon-Germanium Base Heterojunction Bipolar Transistors by Molecular Beam Epitaxy. IEEE Electron Device Lett. 1988, 9 (4), 165–167.

(3)

Weiner, C. How the Transistor Emerged. IEEE Spectr. 1973, 10 (1), 24–33.

(4)

Canham, L. T. Silicon Quantum Wire Array Fabrication by Electrochemical and Chemical Dissolution of Wafers. Appl. Phys. Lett. 1990, 57 (10), 1046.

(5)

Maeda, Y.; Tsukamoto, N.; Yazawa, Y.; Kanemitsu, Y.; Masumoto, Y. Visible Photoluminescence of Ge Microcrystals Embedded in SiO2 Glassy Matrices. Appl. Phys. Lett. 1991, 59 (24), 3168.

(6)

Collings, P. J. Simple Measurement of the Band Gap in Silicon and Germanium. Am. J. Phys 1980, 48 (3), 197–199.

(7)

Maeda, Y. Visible Photoluminescence from Nanocrystallite Ge Embedded in a Glassy Si02 Matrix. Phys. Rev. B 1995, 51 (3).

(8)

Cho, S.; Kang, I. M.; Kim, K. R.; Park, B.; Jr, J. S. H.; Cho, S.; Kang, I. M.; Kim, K. R.; Park, B.; Man Kang, I.; et al. Silicon-Compatible High-Hole-Mobility Transistor with an Undoped Germanium Channel for Low-Power Application. Appl. Phys. Lett. 2013, 103

ACS Paragon Plus Environment

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(22). (9)

Seng, K. H.; Park, M.; Guo, Z. P.; Liu, H. K.; Cho, J. Catalytic Role of Ge in Highly Reversible GeO2/Ge/C Nanocomposite Anode Material for Lithium Batteries. Nano Lett. 2013, 2013 (1), 2–8.

(10)

Lee, D. C.; Pietryga, J. M.; Robel, I.; Werder, D. J.; Schaller, R. D.; Klimov, V. I. Colloidal Synthesis of Infrared-Emitting Germanium Nanocrystals. J. Am. Chem. Soc. 2009, 131 (10), 3436–3437.

(11)

Prabakar, S.; Shiohara, A.; Hanada, S.; Fujioka, K.; Yamamoto, K.; Tilley, R. D. Size Controlled Synthesis of Germanium Nanocrystals by Hydride Reducing Agents and Their Biological Applications. Chem. Mater. 2010, 22 (2), 482–486.

(12)

Lambert, T. N.; Andrews, N. L.; Gerung, H.; Boyle, T. J.; Oliver, J. M.; Wilson, B. S.; Han, S. M. Water-Soluble germanium(0) Nanocrystals: Cell Recognition and nearInfrared Photothermal Conversion Properties. Small 2007, 3 (4), 691–699.

(13)

Choi, W. K.; Chim, W. K.; Heng, C. L.; Teo, L. W.; Ho, V.; Ng, V.; Antoniadis, D. A.; Fitzgerald, E. A. Observation of Memory Effect in Germanium Nanocrystals Embedded in an Amorphous Silicon Oxide Matrix of a Metal-Insulator-Semiconductor Structure. Appl. Phys. Lett. 2002, 80 (11), 2014–2016.

(14)

Kanoun, M.; Busseret, C.; Poncet, A.; Souifi, A.; Baron, T.; Gautier, E. Electronic Properties of Ge Nanocrystals for Non Volatile Memory Applications. Solid. State. Electron. 2006, 50 (7–8), 1310–1314.

(15)

Batra, Y.; Kabiraj, D.; Kanjilal, D. Charge Retention and Optical Properties of Ge Nanocrystals Embedded in GeO2 Matrix. Solid State Commun. 2007, 143 (4–5), 213–216.

(16)

Das, S.; Manna, S.; Singha, R.; Anopchenko, A.; Daldosso, N.; Pavesi, L.; Dhar, A.; Ray, S. K. Light Emission and Floating Gate Memory Characteristics of Germanium Nanocrystals. Phys. status solidi 2011, 208 (3), 635–638.

(17)

Ray, S. K.; Das, K. Luminescence Characteristics of Ge Nanocrystals Embedded in SiO2 Matrix. Opt. Mater. (Amst). 2005, 27 (5), 948–952.

(18)

Seng, K. H.; Park, M.-H.; Guo, Z. P.; Liu, H. K.; Cho, J. Self-Assembled Germanium/carbon Nanostructures as High-Power Anode Material for the Lithium-Ion Battery. Angew. Chem. Int. Ed. Engl. 2012, 51 (23), 5657–5661.

(19)

Riabinina, D.; Durand, C.; Chaker, M.; Rowell, N.; Rosei, F. A Novel Approach to the Synthesis of Photoluminescent Germanium Nanoparticles by Reactive Laser Ablation. Nanotechnology 2006, 17 (9), 2152–2155.

(20)

Giri, P. K. Strain Analysis on Freestanding Germanium Nanocrystals. J. Phys. D. Appl. Phys. 2009, 42 (24), 245402.

(21)

Gorla, C. R. Silicon and Germanium Nanoparticle Formation in an Inductively Coupled Plasma Reactor. J. Vac. Sci. Technol. A Vacuum, Surfaces, Film. 1997, 15 (3), 860.

(22)

Kornowski, A.; Giersig, M.; Vogel, R.; Chemseddine, A.; Weller, H. Nanometer‐sized Colloidal Germanium Particles: Wet‐chemical Synthesis, Laser‐induced Crystallization and Particle Growth. Adv. Mater. 1993, 5 (9), 634–636.

(23)

Chiu, H. W.; Chervin, C. N.; Kauzlarich, S. M. Phase Changes in Ge Nanoparticles. 2005,

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

95616 (19), 4858–4864. (24)

Fok, E.; Shih, M.; Meldrum, A.; Veinot, J. G. C. Preparation of Alkyl-Surface Functionalized Germanium Quantum Dots via Thermally Initiated Hydrogermylation. Chem. Commun. (Camb). 2004, 1 (4), 386–387.

(25)

Taylor, B. R.; Kauzlarich, S. M.; Lee, H. W. H.; Delgado, G. R. Solution Synthesis of Germanium Nanocrystals Demonstrating Quantum Confinement. Chem. Mater. 1998, 10 (1), 22–24.

(26)

Ma, X.; Wu, F.; Kauzlarich, S. M. Alkyl-Terminated Crystalline Ge Nanoparticles Prepared from NaGe: Synthesis, Functionalization and Optical Properties. J. Solid State Chem. 2008, 181 (7), 1628–1633.

(27)

Taylor, B. R.; Kauzlarich, S. M.; Delgado, G. R.; Lee, H. W. H.; Nanocrystals, L. E. S. Solution Synthesis and Characterization of Quantum Confined Ge Nanoparticles. Chem. Mater. 1999, 11 (9), 2493–2500.

(28)

Tanke, R. S.; Kauzlarich, S. M.; Patten, T. E.; Pettigrew, K. a.; Murphy, D. L.; Thompson, M. E.; Lee, H. W. H. Synthesis of Germanium Nanoclusters with Irreversibly Attached Functional Groups: Acetals, Alcohols, Esters, and Polymers. Chem. Mater. 2003, 15 (8), 1682–1689.

(29)

Gerion, D.; Zaitseva, N.; Saw, C.; Casula, M. F.; Fakra, S.; Van Buuren, T.; Galli, G. Solution Synthesis of Germanium Nanocrystals: Success and Open Challenges. Nano Lett. 2004, 4 (4), 597–602.

(30)

Lu, X.; Ziegler, K. J.; Ghezelbash, A.; Johnston, K. P.; Korgel, B. a. Synthesis of Germanium Nanocrystals in High Temperature Supercritical Fluid Solvents. Nano Lett. 2004, 4 (5), 969–974.

(31)

Zaitseva, N.; Dai, Z. R.; Grant, C. D.; Harper, J.; Saw, C. Germanium Nanocrystals Synthesized in High-Boiling-Point Organic Solvents. Chem. Mater. 2007, 19 (21), 5174– 5178.

(32)

Muthuswamy, E.; Iskandar, A. S.; Amador, M. M.; Kauzlarich, S. M. Facile Synthesis of Germanium Nanoparticles with Size Control: Microwave versus Conventional Heating. Chem. Mater. 2012, 25 (8), 1416–1422.

(33)

Vaughn II, D. D.; Schaak, R. E. Synthesis, Properties and Applications of Colloidal Germanium and Germanium-Based Nanomaterials. Chem. Soc. Rev. 2012, 42 (7), 2861– 2879.

(34)

Carolan, D.; Doyle, H. Tuning the Photoluminescence of Germanium Nanocrystals through Surface Bound Functional Groups. Part. Part. Syst. Charact. 2016.

(35)

Hessel, C. M.; Henderson, E. J.; Veinot, J. G. C.; Uni, V.; February, R. V; Re, V.; Recei, M.; August, V. A Molecular Precursor for Nanocrystalline Si - SiO2 Composites and Freestanding Hydride-Surface-Terminated Silicon Nanoparticles. Chem. Mater. 2006, 18, 6139–6146.

(36)

Hessel, C. M.; Henderson, E. J.; Veinot, J. G. C. An Investigation of the Formation and Growth of Oxide-Embedded Silicon Nanocrystals in Hydrogen Silsesquioxane-Derived Nanocomposites. J. Phys. Chem. C 2007, 111, 6956–6961.

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(37)

SI Yang, Z.; Dobbie, A. R.; Cui, K.; Veinot, J. G. C.; Yang, Z.; Dobbie, A. R.; Cui, K.; Veinot, J. G. C. A Convenient Method for Preparing Alkyl-Functionalized Silicon Nanocubes. J. Am. Chem. Soc. 2012, 134 (34), 13958–13961.

(38)

Yang, Z.; Iqbal, M.; Dobbie, A. R.; Veinot, J. G. C. Surface-Induced Alkene Oligomerization : Does Thermal Hydrosilylation Really Lead to Monolayer Protected Silicon Nanocrystals ? 2013.

(39)

Henderson, E. J.; Hessel, C. M.; Veinot, J. G. C. Synthesis and Photoluminescent Properties of Size-Controlled Germanium Nanocrystals from Phenyl TrichlorogermaneDerived Polymers. J. Am. Chem. Soc. 2008, 130 (11), 3624–3632.

(40)

Hoffman, M.; Veinot, J. G. C. Understanding the Formation of Elemental Germanium by Thermolysis of Sol-Gel Derived Organogermanium Oxide Polymers. Chem. Mater. 2012, 24 (7), 1283–1291.

(41)

Jolly, W. L.; Latimer, W. M. The Solubility of Hydrous Germanous Oxide and the Potential of the Germanous Oxide-Germanic Oxid Couple. J. Am. Chem. Soc. 1952, 74, 5751–5752.

(42)

Yang, D. J.; Jolly, W. L.; O’Keefe, A. Conversion of Hydrous germanium(II) Oxide to Germynyl Sesquioxide (HGe)2O3. Inorg. Chem. 1977, 16 (11), 2980–2982.

(43)

Dennis, L. M.; Hulse, R. E. Germanium XXXV. Germanium Monoxide. Germanium Mono Sulfide. J. Am. Chem. Soc. 1930, 52, 3553–3556.

(44)

Gayer, K. H.; Zajicek, O. T. The Solubility of Germanium (II) Hydroxide in Perchloric Acid and Sodium Hydroxide at 25 °C. J. Inorg. Nucl. Chem. 1964, 26, 2123–2125.

(45)

Babich, O. A.; Ghosh, M. C.; Gould, E. S. Preparation of Aqueous Solutions of Hypovalent Germanium; Reactions Involving Germanium-(II) and -(III). Chem. Commun. 2000, No. 11, 907–908.

(46)

Terrey, H.; Everest, D. A.; Terrey, H. Germanous Oxide and Sulphide. J. Chem. Soc. 1950, 2282–2285.

(47)

Rochow, E. G.; Abel, E. W. The Chemistry of Germanium: Tin and Lead; Elsevier, 2014; Vol. 14.

(48)

Chiu, H. W.; Chervin, C. N.; Kauzlarich, S. M. Phase Changes in Ge Nanoparticles. Chem. Mater. 2005, 17 (19), 4858–4864.

(49)

Atuchin, V. V.; Gavrilova, T. A.; Gromilov, S. A.; Kostrovsky, V. G.; Pokrovsky, L. D.; Troitskaia, I. B.; Vemuri, R. S.; Carbajal-Franco, G.; Ramana, C. V. Low-Temperature Chemical Synthesis and Microstructure Analysis of GeO 2Crystals with α-Quartz Structure. Cryst. Growth Des. 2009, 9 (4), 1829–1832.

(50)

Javadi, M.; Yang, Z.; Veinot, J. G. C. Surfactant-Free Synthesis of GeO(2) Nanocrystals with Controlled Morphologies. Chem. Commun. (Camb). 2014, 50 (46), 6101–6104.

(51)

Kuhn, W. K. Zinc Germanium Phosphide by XPS. Surf. Sci. Spectra 1994, 3 (2), 93–99.

(52)

Purkait, T. K.; Swarnakar, A. K.; Glenda, B.; Hegmann, F. A.; Rivard, E.; Veinot, J. G. C. One-Pot Synthesis of Functionalized Germanium Nanocrystals from a Single Source Precursor. Nanoscale 2015, 7 (6), 2241–2244.

(53)

Strahlenchemie, B.; Heights, Y.; Heights, Y.; Schmeisser, D.; Schnell, R. D.; Bogen, A.;

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Himpsel, F. J.; Rieger, D.; Landgren, G.; Morar, J. F. Surface Oxidation States of Germanium. Surf. Sci. 1986, 172 (2), 455–465. (54)

Heath, J. R.; Shiang, J. J.; Alivisatos, a P. Germanium Quantum Dots: Optical Properties and Synthesis. J. Chem. Phys. 1994, 101 (2), 1607–1615.

(55)

Fujii, M.; Hayashi, S.; Yamamoto, K. Growth of Ge Microcrystals in SiO2 Thin Film Matrices. A Raman and Electron Microscopic Study. Japanese J. Appl. Physics, Part 1 Regul. Pap. Short Notes Rev. Pap. 1991, 30, 687–694.

(56)

Karavanskii, V. a.; Lomov, a. a.; Sutyrin, a. G.; Bushuev, V. a.; Loikho, N. N.; Melnik, N. N.; Zavaritskaya, T. N.; Bayliss, S. Raman and X-Ray Studies of Nanocrystals in Porous Stain-Etched Germanium. Thin Solid Films 2003, 437 (1-2), 290–296.

(57)

Gayer, K. H.; Zajicek, O. T. The Solubility of Germanium (IV) Oxide in Aqueous NaOH Solutions at 25°C. J. Inorg. Nucl. Chem. 1964, 26 (6), 951–954.

(58)

Yang, Z.; Gonzalez, C. M.; Purkait, T. K.; Iqbal, M.; Meldrum, A.; Veinot, J. G. C. Radical Initiated Hydrosilylation on Silicon Nanocrystal Surfaces: An Evaluation of Functional Group Tolerance and Mechanistic Study. Langmuir 2015, 31 (38), 10540– 10548.

(59)

Fuchs, R.; Moore, L.; Miles, D.; Gilmak, H. Some Symmetrical Tetraalkyl- and Tetraaralkyl-Germanes. J. Org. Chem. 1956, 1113–1117.

(60)

Sun, X. H.; Didychuk, C.; Sham, T. K.; Wong, N. B. Germanium Nanowires: Synthesis, Morphology and Local Structure Studies. Nanotechnology 2006, 17 (12), 2925.

(61)

Buriak, J. M. Organometallic Chemistry on Silicon and Germanium Surface. Chem. Rev. 2002, 102 (5), 1271.

(62)

Bentlohner, M. M.; Waibel, M.; Zeller, P.; Sarkar, K.; Müller-Buschbaum, P.; FattakhovaRohlfing, D.; Fässler, T. F. Zintl Clusters as Wet-Chemical Precursors for Germanium Nanomorphologies with Tunable Composition. Angew. Chemie Int. Ed. 2016, 55 (7), 2441–2445.

(63)

Guloy, A. M.; Ramlau, R.; Tang, Z.; Schnelle, W.; Baitinger, M.; Grin, Y. A Guest-Free Germanium Clathrate. Nature 2006, 443 (7109), 320–323.

(64)

Cheng, X.; Lowe, S. B.; Ciampi, S.; Magenau, A.; Gaus, K.; Reece, P. J.; Gooding, J. J. Versatile “Click Chemistry” Approach to Functionalizing Silicon Quantum Dots: Applications toward Fluorescent Cellular Imaging. Langmuir 2014, 30 (18), 5209–5216.

(65)

Purkait, T. K.; Iqbal, M.; Wahl, M. H.; Gottschling, K.; Gonzalez, C. M.; Islam, M. A.; Veinot, J. G. C. Borane-Catalyzed Room-Temperature Hydrosilylation of Alkenes/Alkynes on Silicon Nanocrystal Surfaces. J. Am. Chem. Soc. 2014, 136 (52), 17914–17917.

(66)

Clayden, J.; Greeves, N.; Warren, S. G. Organic Chemistry.; Oxford ; New York : Oxford University Press, 2012., 2012.

(67)

Hua, F.; Swihart, M. T.; Ruckenstein, E. Efficient Surface Grafting of Luminescent Silicon Quantum Dots by Photoinitiated Hydrosilylation. Langmuir 2005, 21 (13), 6054– 6062.

(68)

Gottlieb, H. E.; Kotlyar, V.; Nudelman, A. NMR Chemical Shifts of Common Laboratory

ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Solvents as Trace Impurities. J. Org. Chem. 1997, 62 (21), 7512–7515. (69)

Fulmer, G. R.; Miller, A. J. M.; Sherden, N. H.; Gottlieb, H. E.; Nudelman, A.; Stoltz, B. M.; Bercaw, J. E.; Goldberg, K. I. NMR Chemical Shifts of Trace Impurities: Common Laboratory Solvents, Organics, and Gases in Deuterated Solvents Relevant to the Organometallic Chemist. Organometallics 2010, 29 (9), 2176–2179.

(70)

Ruddy, D. A.; Johnson, J. C.; Smith, E. R.; Neale, N. R.; Near-, S. S.; Ruddy, D. A.; Johnson, J. C.; Smith, E. R.; Neale, N. R. Size and Bandgap Control in the Solution-Phase Synthesis of Near-Infrared-Emitting Germanium Nanocrystals. ACS Nano 2010, 4 (12), 7459–7466.

(71)

Maeda, Y.; Tsukamoto, N.; Yazawa, Y.; Kanemitsu, Y.; Masumoto, Y. Visible Photoluminescence of Ge Microcrystals Embedded in SiO2 Glassy Matrices. Appl. Phys. Lett. 1991, 59 (24), 3168–3170.

(72)

Dutta, A. K. Visible Photoluminescence from Ge Nanocrystal Embedded into a SiO2 Matrix Fabricated by Atmospheric Pressure Chemical Vapor Deposition. Appl. Phys. Lett. 1996, 68 (9).

(73)

Takeoka, S.; Fujii, M.; Hayashi, S.; Yamamoto, K. Size-Dependent near-Infrared Photoluminescence from Ge Nanocrystals Embedded in SiO 2 Matrices. Phys. Rev. B 1998, 58 (12), 7921.

(74)

Zacharias, M.; Fauchet, P. M. Blue Luminescence in Films Containing Ge and GeO2 Nanocrystals: The Role of Defects. Appl. Phys. Lett. 1997, 71, 380–382.

(75)

Min, K. S.; Shcheglov, K. V; Yang, C. M.; Atwater, H. A.; Brongersma, M. L.; Polman, A. The Role of Quantum‐confined Excitons vs Defects in the Visible Luminescence of SiO2 Films Containing Ge Nanocrystals. Appl. Phys. Lett. 1996, 68 (18), 2511–2513.

(76)

Weber, J. R.; Janotti, A.; Van de Walle, C. G. Defects in Germanium. In Photonics and Electronics with Germanium; Wiley-VCH Verlag GmbH & Co. KGaA, 2015; pp 1–23.

(77)

Popescu, D. P.; Eliseev, P. G.; Stintz, A.; Malloy, K. J. Temperature Dependence of the Photoluminescence Emission from InAs Quantum Dots in a Strained Ga0. 85In0. 15As Quantum Well. Semicond. Sci. Technol. 2003, 19 (1), 33.

(78)

Nardis, S.; Mandoj, F.; Paolesse, R.; Fronczek, F. R.; Smith, K. M.; Prodi, L.; Montalti, M.; Battistini, G. Synthesis and Functionalization of Germanium Triphenylcorrolate: The First Example of a Partially Brominated Corrole. Eur. J. Inorg. Chem. 2007, 2007 (16), 2345–2352.

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC/Graphical Abstract 576x186mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Scheme 1. Preparation of dodecane functionalized GeNCs. (1), Thermal processing of ‘Ge(OH)2’ at T = 400 °C in Ar, (2), Liberation of H-GeNCs via HF etching, (3), Functionalization/surface modification of GeNCs through thermally-activated, radical-initiated, or borane-catalyzed hydrogermylation. 1019x276mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1 Characterization of ‘Ge(OH)2’ before and after thermal processing at 400oC (for 1 h under Ar) (a) X-ray powder diffraction patterns. Standard reflections of crystalline Ge are provided.49,50 Rutile GeO2 reflections are indicated using a *. (b) XP spectra of the Ge 3d spectral region. (c) Raman spectra. 908x391mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 2 Characterization of GeNCs obtained from the thermal treatment of ‘Ge(OH)2’ at 400 °C for 1 h in a flowing Ar atmosphere. (a) FT-IR spectra of hydride-terminated (black trace) and dodecyl-GeNCs obtained from thermal hydrogermylation (red trace). (b) Bright-field TEM micrograph of dodecyl-GeNCs. (c) and (d) EDX and Raman spectra of dodecyl-GeNCs obtained from thermal hydrogemylation. 1400x527mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. Nanoassisted laser desorption ionization mass spectra of dodecyl-GeNCs functionalized using a) thermally-activated (inset: higher m/z showing dodecene repeat units) b) AIBN and c) BP radical-initiated, d) borane-catalyzed hydrogermylation. Inset spectra are higher magnification from larger m/z. 814x688mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 4. 1H NMR spectrum of dodecyl-GeNCs obtained from thermally activated hydrogermylation in CDCl3 containing 0.01% (v/v) TMS. Integration of dodecyl methyl protons (3H) to TMS protons (12H) are shown. Methyl and chain methylene protons denoted by a and b, respectively. Residual solvent impurities (H2O or HOD at 1.5 ppm and toluene CH3 at 2.34 ppm)64,65 are denoted by (∗). 489x398mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 28 of 28